Sei sulla pagina 1di 7

ENERGY LEVELS AND TRANSITIONS IN ATOMS ENERGY LEVELS IN ATOM

Generally, atoms consist of a positive nucleus surrounded by a cloud of negative electrons. Each electron in the cloud may possess only very specific amounts of energy. The total electronic energy of the atom is the sum of the energies of all its individual electrons. If the electronic energy of the atom is such that it contains only the minimum allowed energy, the atom is said to be in the "ground state." If the total energy content of the atom is greater than the ground state energy, the atom is said to be in an "excited state." Figure is a partial energy level diagram for a mercury atom. The ground state is the energy level denoted as E . !hen in this state, the atom has an electronic energy labelled as "ero. This "ero does not mean that the atom contains no energy but, rather, that it contains its minimum allowable energy and that no electronic energy can be removed from it.

Fig. 1 of mercury #ome energy levels$ The higher energy levels indicated %E&, E', etc.( indicate specific amounts of energy that the atom may contain. Each of these levels corresponds to a particular configuration for the electrons around the nucleus of the atom. In general, an electronic configuration which, on the average, has its electrons further removed from the nucleus than others, will possess a higher energy state, hence a higher atomic energy level in Figure . ) single atom may occupy only one of these energy levels at any one instant. In order to move from one energy level to another, the atom must gain or lose an amount of energy exactly e*ual to the energy difference between the two levels. #uch a change in energy level such+as that shown by the arrows in Figure from E, to E& or E to E-+is called an "atomic transition," and this change may occur in several ways. )n atom in an excited state+that is, any state above the ground state+will not remain there indefinitely. )toms tend to release their excess energy and return to the ground state or by a series of transitions to successively lower energy levels, ending at the ground state. The atomic lifetime of a particular energy state is the time re*uired for half of the atoms initially in that state to ma.e a downward transition without benefit of outside influence %such as stimulated emission(. For example

if / & mercury atoms were initially in energy state E0 of Figure , only , x / atoms would remain in that state after a time interval e*ual to the atomic lifetime of that state. The atomic lifetime, therefore, is a measure of the rate at which atoms leave a given energy level by releasing some of their energy. The average atomic lifetime is about /1- seconds, but there are large variations. )tomic lifetimes may be as short as /1 sec or as long as /1& sec. Energy states having atomic lifetimes of /10 sec or longer are called "metastable states."

SPONTANEOUS EMISSION OF A P OTON


2ne way for an atom to ma.e a transition from a particular energy level to a lower energy state is by a process .nown as "spontaneous emission." This type of emission occurs when the atom releases a photon possessing an energy e*uivalent to the energy difference between the two energy levels in *uestion. The photon is emitted without benefit of an external stimulus and travels away from the point of emission in a random direction. The fre*uency and wavelength of the photon are, of course, a function of the energy of the emitted photon.

RA!IATIONLESS TRANSITIONS
In some cases, atoms ma.e downward transitions without releasing a photon. !hen this occurs, the energy released by the atom must be carried away in some form other than emitted electromagnetic radiation. #uch a transition is called a ""a#iatio$less t"a$sitio$." In gases, radiationless transitions occur when an excited atom collides with an atom in some lower energy state. ) portion of the energy of the more energetic atom is transferred to the less energetic atom during the collision. 2ne atom loses a certain amount of energy while the other atom absorbs the energy lost by the more energetic atom. In this manner, the energy is released by an atom, ma.ing the downward transition without the production of a photon. In solids, radiationless transitions often account for a temperature increase within the material. The energy released by an atom in a downward transition appears as other %e.g., vibrational( energy in the area surrounding the atom3s e*uilibrium position in the solid. This vibrational energy increases the thermal energy in the solid and, thus, raises its temperature.

A%SORPTION OF A P OTON %Y AN ATOM


4nder certain conditions, atoms can absorb the same wavelengths of light that they emit through spontaneous emission. ) case in point is illustrated in Figure &.

Fig. & )bsorption of light )n atom initially in energy state E is struc. by a photon of energy Ep 5 E' 1 E . The photon ceases to exist, and the atom ma.es an upward transition to level E'. The photon3s energy now is contained in the excited atom in the E' energy level. For this type of photon absorption by an atom to occur, two conditions must be satisfied6

The energy of the incoming photon must be e*uivalent to the energy difference between the two energy levels in *uestion. The atom absorbing the photon must be in the lower of the two energy levels.

Thus, an atom in level E in Figure & could absorb a photon of energy %E& 1 El(, %E' 1 E (, or %E7 1 E (, but it could not absorb a photon of energy %E7 1 E&( or %E' 1 E&(.

STIMULATE! EMISSION
Figure ' illustrates the stimulated emission process that produces laser light. The conditions necessary for stimulated emission to occur are the same as those for absorption, except that the emitting atom must be in the upper of the two energy states involved. In the case of stimulated emission, the incident photon stimulates the atom to release a photon sooner than it would have in the absence of an external stimulus, as in spontaneous emission. In this case, the photon released by the stimulated atom has the same energy, fre*uency, wavelength, phase and direction of travel as the incident stimulating photon. The photon emitted by the stimulated atom is also in phase with the incident photon, and the energies of both photons are added together in the resultant beam. This, or course, is the process at the heart of laser action.

Fig. ' Stim(late# emissio$ of lig)t Emissio$ S*ect"a fo" Atomic Gases To observe an emission spectrum, one creates an electric discharge in a closed, transparent tube which contains the gas, thereby causing the gas to glow or fluoresce. This fluorescent light is sent through a well8defined, narrow entrance slit of a spectrometer. The spectrometer, with the help of prisms or gratings, forms separate images of the entrance slit on a photographic plate, depending on the different wavelengths of fluorescent light passing through the slit. ) typical setup designed to observe the emission spectrum for neon gas, for example, is shown in Figure 7a. The line images formed on the plate are shown in Figure 7b. In Figure 7c, the same line images are shown, appropriately coloured according to their actual wavelengths, 9ust as they would be seen by the na.ed eye.

The formation of emission spectra and their relationship to the energy levels characteristic of the atomic gases is not hard to understand. The energy provided by the electric discharge is absorbed by the atoms in the gas, causing them to be raised to higher energy levels+levels such as E&, E', E7 and so on in Figure . 2nce the atoms reach the various higher energy levels, they begin to fall bac. to lower levels via the process of spontaneous emission. %It is the spontaneous emission that we see as the tube glows or fluoresces.( )s the atoms fall bac.+say from level E- to E:, or E: to E7, or E7 to E , or directly from E- to E +they emit photons with energies e*ual to E- 1 E:, E: 1 E7, E7 1 E , E- 1 E and so on. #ince we .now that l 5 hc;E, we see that, for each transition, a photon of different wavelength is created. !hen these emitted photons+arising from the downward energy transitions in the excited gas+are sent through the entrance slit of a spectrometer, they form separate, distinct images of the spectrometer slit on the photographic plate. If we have .nown wavelengths of calibration lines on the plate, we can calculate the wavelength of the test8gas lines, and wor. bac. to get the energy differences and finally the various energy levels for the test gas. In this way we establish the energy level diagrams+the emission spectra+which characteri"e the various atomic gases such as hydrogen, helium, neon, etc. ) closer loo. at the line images on the plate shows that they are not all e*ually bright. #ome are lighter, some are dar.er. )n analysis of the "intensity" of each line formed on the plate gives us important information about the probability for a transition between the various energy levels. For example, if a certain line, say that for the transition from E- to E , is much more intense than the line for the transition from E- to E7, then the transition probability for spontaneous emission from E- to E is higher than that from E- to E7. Abso"*tio$ S*ect"a fo" Atomic Gases The process for observing absorption spectra is similar to that for observing emission spectra. In this case one places the atomic gas under study in a transparent, cylindrical container and passes a collimated, white8light beam through the container. The light beam emerging from the cylinder is then

analysed by a spectrometer, with the absorption lines showing up on a photographic plate. The lab setup is shown schematically in Figure ,.

In absorption spectra, the photographic plate would be uniformly exposed %dar.ened( due to the white light, except at positions where the particular wavelengths in the white light are absorbed by the gas. For those absorbed wavelengths, no light reaches the plate, and at those positions the plate "registers" line images for the "absent" light. #ince "white" light contains all wavelengths in the visible spectrum, a continuum from 7// nm to :// nm, the beam incident on the atomic gas contains a continuum of photon energies, from .:: e< %for l 5 :// nm( to '. / e< %for l 5 7// nm(. For each photon in the beam, for which a match between the photon energy and an energy level difference in the atomic gas exists, the photon is absorbed and disappears from the beam. )ll other photons continue on through the spectrometer and expose the photographic plate. The absorbed photons, with their tell8tale wavelengths, never reach the plate, so the plate remains unexposed at those particular wavelengths+resulting in white lines on a dar. bac.ground. !hen the plate is developed and a positive print is made from the negative, the spectrum appears with blac. lines on a white bac.ground. Identifying the wavelengths of these lines and wor.ing bac. to establish the energy level differences for the atomic gas, we develop the energy level diagram characteristic of that gas, 9ust as was done with emission spectra.

Fig. + Li$e s*ect"(m of a mo$atomic atom )gain, the intensity of the various lines on the photographic plate provides evidence for the transition probability. )n intensity analysis of the absorption lines8performed with an optical instrument called a densitometer8 yields a trace such as that shown in Figure 0. =ere the line emission intensity is plotted against wavelength for the line spectra. In the trace, the higher the pea.s, the more probable the transition and the shorter the lifetime of the energy state involved. #imilarly, the lower the pea.s, the less probable the transition and the longer the lifetime of the state involved.

EMISSION AN! A%SORPTION SPE,TRA ATOMI, GASES Figure 7 represents the typical emission spectrum of a monatomic gas, such as helium or neon. This type of spectrum is produced by an electrical discharge passed through a gas sample contained at low pressure. Each of the lines in the emission spectrum is produced by a single atomic transition, and the intensity of each line produced is dependent upon the probability of atoms ma.ing that particular transition. #tronger lines are the result of the most probable transitions from energy states having short atomic lifetimes. !ea.er lines are produced by transitions from states that have long atomic lifetimes or by low8probability transitions that compete with the higher8probability transitions. Each of the lines within the line spectrum %Figure 0( actually consists of a narrow "a$ge of wavelengths %Figure :(, which is the result of a phenomenon called "!o**le" b"oa#e$i$g."

Fig. - .i#t) of a s*ect"al li$e )s illustrated in Figure -, stationary atoms emit light of a wavelength l/, corresponding to the center of the spectral line, although moving atoms may emit slightly different wavelengths because of their motion. If a moving atom emits a photon in the same direction of travel as the atom, the wavelength %l ( will be slightly shorter %fre*uency higher( than the wavelength emitted by a stationary atom. In contrast, if the moving atom emits a photon traveling in a direction opposite to the atom3s motion, the wavelength %l&( will be longer %fre*uency lower(. Thus, the additive result for a large collection of atoms is the >oppler broadened line of Figure ,. The line center energy corresponds to the exact difference in energies of the two electronic states involved in the transition for stationary atoms or atoms moving in the same direction at the same velocity.

Fig. / T)e !o**le" effect.

SOLIDS
Figure ? represents the trace of an absorption spectrum for a solid material. 4nli.e the comparatively narrow spi.es of a gaseous spectrum, this spectrum consists of broad, irregularly8spi.ed regions

called absorption bands. This difference in spectral character is due to the fact that the energy levels of an atom bound in a solid shift slightly in the ever8present local electric and magnetic fields. Each atom bound within a material produces electric and magnetic fields as a characteristic of the atom3s nature@ conse*uently, when large numbers of atoms are crowded close together in a solid, the energy levels of each atom shift because of the fields produced by all its neighbours. This wholesale shifting of energy levels broadens all the spectral lines. )reas containing a closely8spaced group of strong lines appear as an absorption band. In crystalline solids such as Ad6B)G %Ad atoms in yttrium aluminium garnet(, these absorption bands consist of groups of sharp8edged lines. In solids such as Ad6glass which lac. an ordered crystal structure, the bands are broader and less distinct. The definitions of these absorption bands are important in determining the manner of optical exciting solid8state lasers.

Fig. 0 Abso"*tio$ s*ect"(m of a soli# Resol(tio$ 1.i#t)2 Li$es S*ect"a )tomic spectral lines have finite widths with factors to line broadening due to6 C Aatural Droadening 8 The lifetime of the excited states lead to uncertainty leading to broadening due to shorter excited state lifetimes. Eifetimes of /8 - s lead to width of /8, nm. C Follisional Droadening 8)lso referred to as P"ess("e %"oa#e$i$gis the result of collision of the excited state leads to shorter lifetimes and broadening of the spectral lines. C >oppler Droadening 8 !hen molecules are moving towards a detector or away from a detector the fre*uency will be offset by the net speed the radiation hits the detector. This is also .nown as the >oppler effect and the true fre*uency will ether be red shifted %if the chemical is moving away from the detector( or blue shifted %if the chemical is moving towards the detector(

Potrebbero piacerti anche